首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 339 毫秒
1.
We synthesized complexones based on tetrapyridylporphin (I), namely, tetra(pyridinium-4-N-carboxymethylene) porphin tetrachloride (II), tetra(pyridinium-4-N-carboxymethylene)porphin tetrabromide (III), complete ethyl ester of (pyridinium-4-N-carboxymethylene)porphin tetrachloride (IV), and tetra(pyridinium- 4-N-methyl)porphin tetraiodide (V). Data on the electronic absorption spectra of the compounds, the enthalpies of their solution in water, and the enthalpies of complex formation with copper(II) and zinc acetates were obtained, and the acid properties of the ligands and their stability under thermooxidative destruction conditions were studied. The conclusion was drawn that the enthalpy of solution of complexone-porphyrins in water was determined by the strength of their crystal lattices and, when different anions (Cl?, Br?, and I?) were present, by the difference of the enthalpies of their hydration. As distinct from Cu2+, complex formation between Zn2+ and porphyrin ligands occurred with sensible energy expenditures likely caused by the electronic inconsistency between the zinc cation and porphyrin ligands. The stability of water-soluble porphyrins to thermooxidative destruction was limited by temperatures of 200–260°C.  相似文献   

2.
Kim BH  Lee do N  Park HJ  Min JH  Jun YM  Park SJ  Lee WY 《Talanta》2004,62(3):595-602
A series of o-phenanthroline-substituted ruthenium(II) complexes containing 2,2′-dipyridyl, 2-(2-pyridyl)benzimidazole, 2-(2-pyridyl)-N-methylbenzimidazole, 4-carboxymethyl-4′-methyl-2,2′-dipyridyl, and/or 4,4′-dimethyl-2,2′-dipyridyl ligands were synthesized and examined as potent electrochemiluminescent (ECL) materials. The characteristics of these complexes, regarding their electrochemical redox potentials and relative ECL intensities for tripropylamine were studied. As found in a 2,2′-bipyridyl-substituted ruthenium(II) complexes, a good correlation between the observed ECL intensity and the donor ability of α-diimine ligands was observed, i.e., the ECL intensity of the Ru(II) complex decreased with an increase in the ligand donor ability. The ECL efficiency increased as the number of substitutions of o-phenanthroline (o-phen) to metal complexes increased.  相似文献   

3.
Palladium(II) complexes containing di-(2-pyridyl)-N-methylimine (1), di-(2-pyridyl)methanol (2) and di-(2-pyridyl)methyl-N,N-diethyldithiocarbamate (4) ligands were synthesized and characterized by 1H and 13C NMR in solution, IR and X-ray single crystal diffraction. Crystal structures of cis-dichloro[di-(2-pyridyl)-N-methylimine]palladium(II) (5), cis-dichloro[di-(2-pyridyl)methanol]palladium(II) (6) and cis-dichloro[di-(2-pyridyl)methyl-N,N-diethyldithiocarbamate]palladium(II) (7) showed a bidentate coordination mode of the di-(2-pyridyl)methane derivatives 1, 2 and 4. In these complexes is observed the formation of a five-membered chelate ring with the iminic ligand 1 and six-membered chelate rings with the pyridinic ligands 2 and 4. In all complexes the palladium atom displays a distorted square planar geometry.  相似文献   

4.
The synthesis and reactions of 1-(N-benzylamino)-1-(2-pyridyl)- and 1-(N-benzylamino)-1-(4-pyridyl)-methyldiphenylphosphine oxides are described. It was found that these compounds were exceptionally easy to cleave in aqueous sulfuric acid solutions to form diphenylphosphinic acid and the corresponding N-(pyridylmethyl)-benzylamines. The structure of a single diastereoisomer, that is, the (R)-(+)-1-[N-(α-methylbenzylamino)]-1-(4-pyridyl)-(S)-methyldiphenylphosphine oxide was determined by X-ray crystallography. The acidic alcoholysis of the selected model chiral pyridine aminophosphine oxides was investigated by means of 31P NMR spectroscopy. The cleavage kinetics were also studied. On the basis of the obtained results, a mechanism of the cleavage was formulated.  相似文献   

5.
Excess enthalpies for binary mixtures (S-fenchone + ethanol/benzene/cyclohexane/carbon tetrachloride) were measured over the whole concentration at T = 298.15 K. The experimental results were compared with the values obtained from the UNIFAC, COSMO-RS and regular solution theory. Excess enthalpies of binary mixtures of R-fenchone and S-fenchone in ethanol, benzene, and cyclohexane solution at different specified mole fractions of fenchone have been measured under the same conditions. With the decreasing of the specified mole fraction of fenchone in different solutions, the excess enthalpies of mixing of chiral orientated solutions increased and became close to zero. Results were compared with those of chiral limonene in ethanol solution. Pair interaction energies were also investigated.  相似文献   

6.
Kinetics of the incorporation of mercury(II) ion in tetra (p-trimethylammoniumphenyl)porphine have been investigated in aqueous solution at 30.0°C and 0.2 M (NaNO3) ionic strength. The reaction was found to be first order each in mercury(II) and the porphyrin. The forward (formation) and the reverse (dissociation) rate constants were found to be 1.9 ± 0.2 × 103 M?1 s?1 and 7 ± 2 × 106 M?1 s?1, respectively. Kinetics of zinc(II) incorporation in tetra(p-trimethylammoniumphenyl)porphine catalyzed by mercury(II) were also investigated. This catalysis is explained in terms of steady-state formation of mono mercury(II) porphyrin followed by zinc(II) displacement of mercury(II) ion from the porphyrin. Such a mechanism also illustrates the importance of porphyrin core deformation to metal incorporation.  相似文献   

7.
Heating a neat 1:2 mixture of 2-picolylamine and 2-cyanopyridine followed by treatment of the resultant red gummy substance with aqueous KOH resulted in the isolation of 2,4,5-tris(2-pyridyl)imidazole (1a) as the major product and N-(3-(2-pyridyl)imidazo[1,5-a]pyridine)picolinamidine (2a) in small amounts. Similarly, by using 3-picolylamine, 2,4,-bis(2-pyridyl)-5-(3-pyridyl)imidazole (1b) and N-(3-(3-pyridyl)imidazo[1,5-a]pyridine)picolinamidine (2b) were isolated, and by using 4-picolylamine, 2,4,-bis(2-pyridyl)-5-(4-pyridyl)imidazole (1c) and N-(3-(4-pyridyl)imidazo[1,5-a]pyridine)picolinamidine (2c) were isolated. The plausible mechanism of the formation of 1a-c and 2a-c is delineated.  相似文献   

8.
Two model dipeptides, N-tertiobutoxycarbonylsarcosine N′-methylamide (BSMA) and N-tertiobutoxycarbonylsarcosine N′,N′-dimethylamide (BSDA) are investigated by FT-IR spectrometry. The conformation of BSMA is very sensitive to the environment. In solvents of weak polarity (carbon tetrachloride, cyclohexane), BSMA accomodates the extended and seven-membered ring conformation, but in 1,2-dichloroethane, the C7 conformers are greatly destabilized. Hydrogen bonding between BSMA or BSDA and phenols is studied in carbon tetrachloride. The thermodynamic data (equilibrium constants and enthalpies of complex formation) show that the BSMA complexes are stronger than the BSDA complexes. The spectroscopic data suggest that for BSMA, complex formation occurs at the O atom of the amide function while for BSDA about 50% of the complexes are formed on the O atom of the urethane group. The differences between the two sarcosine dipeptides are discussed in terms of cooperative and steric effects. It can be concluded that the global polarity of the medium exerts a greater influence on the conformation of the C7 dipeptides than the specific interactions taking place on a given site of the molecule.  相似文献   

9.
Agata Bia?ońska 《Tetrahedron》2008,64(41):9771-9779
1-(3-Bromopropyl)tetrazole, 2-(3-bromopropyl)tetrazole, 1-(4-bromobutyl)tetrazole, and 2-(4-bromobutyl)tetrazole were synthesized with the aim to prepare flexible bitopic ligands contaning 1- or 2-substituted tetrazole ring linked through 1,3-propylene or 1,4-butylene spacer with pyridylazole or azole unit. Twenty-six novel ligands i.e., α-(pyridylazolyl)-ω-(tetrazolyl)alkanes, α-(tetrazolyl)-ω-(1,2,3-triazolyl)alkanes, and α-(tetrazol-1-yl)-ω-(tetrazol-2-yl)alkanes were prepared by an alkylation of sodium salts of 5-(2-pyridyl)tetrazole, 3-(2-pyridyl)-1,2,4-triazole, 3-(2-pyridyl)pyrazole, 1,2,3-triazole, and 1,2,3,4-tetrazole with N-(ω-bromoalkyl)tetrazoles. An alkylation of 5-(2-pyridyl)tetrazole, 1,2,3,4-tetrazole, and 1,2,3-triazole afforded both N1- and N2-regioisomer whereas in the case of 3-(2-pyridyl)-1,2,4-triazole and 3-(2-pyridyl)pyrazole only N1 isomers were isolated. The positions of alkylation were confirmed by X-ray diffraction studies of 1-(5-(2-pyridyl)tetrazol-2-yl)-4-(tetrazol-1-yl)butane, 1-(3-(2-pyridyl)-1,2,4-triazol-1-yl)-4-(tetrazol-2-yl)butane, 1-(3-(2-pyridyl)pyrazol-1-yl)-4-(tetrazol-1-yl)butane, and 1-(tetrazol-1-yl)-4-(1,2,3-triazol-1-yl)butane. Preliminary investigations of magnetic properties of iron(II) complex with 1-(3-(2-pyridyl)-1,2,4-triazol-1-yl)-4-(tetrazol-1-yl)butane revealed that obtained product exhibit thermally induced spin transition accompanied by the thermochromic effect.  相似文献   

10.
The electrochemical and photochemical reduction of the fully methylated derivative of gold meso-tetrakis-(4-pyridyl)porphine (AuTMPyP) in homogeneous solution gives not only a π-radical anion but also its successive product, phlorin, by disproportionation. The electrostatic fixation of AuTMPyP in a Nafion matrix inhibits the latter undesired reaction against effective charge separation, which is explained by the diffusion constant lower by 8 order of magnitude during electrolytic reduction of AuTMPyP compared to that in homogeneous aqueous solution.  相似文献   

11.
Solution and mixing enthalpies for the orthophosphoric acid (H3PO4)-N,N-dimethylformamide (DMF) system were measured over the whole concentration range at 25 °C. The standard value of solution enthalpy of phosphoric acid in DMF and the standard transference enthalpy of H3PO4 from water to DMF were calculated. The mixing enthalpy concentration dependence permitted making assumptions on complex formation in the system under investigation.  相似文献   

12.
A series of cationic porphyrins with 1-4 positive charges are studied: mono(N-methyl-4-pyridyl)triphenylporphine chloride [Mono], cis(N-methyl-4-pyridyl)diphenylporphine chloride [Cis], tri(N-methyl-4-pyridyl)monophenylporphine chloride [Tri] and tetra(N-methyl-4-pyridyl)porphine chloride [Tetra]. Their photophysical properties are measured in small unilamellar vesicles and compared with those in homogeneous solution. Liposomes of L-alpha-dimyristoyl-phosphatidylcholine (100 nm diameter) and L-alpha-dipalmitoyl-phosphatidylcholine (50 nm diameter) in phosphate-buffered saline (pH = 7.4) or D2O 0.15 M NaCl were used. The effect of the medium microheterogeinity is discussed. The triplet quantum yields in liposomes for all the porphyrins are about 0.7, similar to the value obtained for Tetra in aqueous media. The singlet molecular oxygen quantum yields for the hydrophilic compounds Tri and Tetra are greater than those of the hydrophobic ones, Mono and Cis. Also, association constants (KL) of the dyes to liposomes and their localization within the membranes are determined from fluorescence and fluorescence polarization measurements, respectively. KL values are in the range of 10(4)-10(5) M-1 for all the compounds, indicating that hydrophobic and coulombic interactions between porphyrins and liposomes are responsible for the dye association. Fluorescence polarization experiments indicate that Mono and Cis can penetrate into the lipidic phase, and that Tri and Tetra are located near the polar heads of the lipidic molecules.  相似文献   

13.
An ortho-dimethyl substituted meso-tetrakisarylporphyrin prefunctionalized with triflate groups was prepared in good yield from an accessible 2,6-dimethyl-4-(triflyloxy)benzaldehyde. This porphyrin is an interesting building block, which could directly be engaged in Suzuki cross-coupling reactions, to be further tetra functionalized with 3-pyridyl ligands in yields equal or above 85%. The porphyrin core of the various compounds bearing four remote coordination sites was metalated with zinc(II). The molecular structures of the starting triflate porphyrin derivative and of the zinc(II) porphyrin substituted with four 3-pyridyl groups bearing a protected alcohol were determined using X-ray crystallography.  相似文献   

14.
Bis-monodentate ligands, such as bis(4-pyridyl) derivatives and bis(4-pyridyl-N-oxide), are able to generate polymetallic coordination networks with interesting supramolecular solid-state architectures. This review is devoted to high-dimensionality systems, which are extended by combining two or three organizing forces: metal-coordination, hydrogen bonds and π–π stacking interactions. A special emphasis is given to the following molecules, which play the role of linkers and spacers in the construction of extended frameworks: 4,4′-bipyridine, 1,2-bis(4-pyridyl)ethane, trans-1,2-bis(4-pyridyl)ethene, trans-4,4′-azo-pyridine, 4,4′-bipyridyl-N,N′-dioxide.  相似文献   

15.
An improved and efficient entry to highly functionalized β-(2-pyridyl)- and β-(4-pyridyl)alanines and the corresponding 1,4-dihydro and N-oxide derivatives has been developed by one-pot thermal Hantzsch-type cyclocondensation of aldehyde-ketoester-enamine systems in which one of the reagents (aldehyde or ketoester) was carrying the unmasked but protected chiral glycinyl moiety. Thus coupling N-Boc-O-benzyl aspartate β-aldehyde, acetoacetate and aminocrotonate esters afforded tetrasubstituted β-(4-dihydropyridyl)alanines (75% yield). One of these products was almost quantitatively transformed into the β-(4-pyridyl)alanine derivative which in turn was oxidized to the corresponding N-oxide. Each of these enantiomerically pure (Mosher's amide analysis) heterocyclic α-amino acids was incorporated into a tripeptide by coupling with (S)-phenylalanine. In a similar way tetrasubstituted β-(2-dihydropyridyl)alanine, β-(2-pyridyl)alanine and β-(1-oxido-2-pyridyl)alanine were prepared via Hantzsch cyclocondensation reaction using benzaldehyde, aminocrotonate, and acetoacetate carrying the N-Boc-O-benzyl glycinate moiety. It was shown that the work up of the reaction mixtures derived from the cyclocondensation and oxidation reactions can be carried out by the use of polymer supported reagents and sequestrants thus allowing the isolation of the products in high purity without any chromatography.  相似文献   

16.
Reactions of 4-aryl-2-hydroxy-4-oxobut-2-enoic acids N-(2-pyridyl)amides with diazomethane, diazoethane, diaryldiazomethanes, and diazofluorene lead to the formation of 2-alkoxy-4-aryl-4-oxobut-2-enoic acids N-(2-pyridyl)amides, 3-aroyl-5-methylpyrazole-4-carboxylic acids N-(2-pyridyl)amides, and 3-alkoxy-3-(2-aryl-2-oxoethyl)-2,3-dihydro-2-oxoimidazo[1,2-a]pyridines. The composition and structure of compounds obtained depend on the nucleophilic nature of the diazo compound and on the character of substituents in the aryl and pyridine parts of the substrate.  相似文献   

17.
《中国化学快报》2022,33(8):4101-4106
Fluorescence (FL) imaging guided photodynamic therapy (PDT) is becoming highly desirable for personalized therapy and precision medicine. In this study, fluorescent polymer nanoparticles TCPP@PEI/PGA were facilely synthesized through electrostatic interaction-mediated self-assembly of porphyrins tetra(4-carboxyphenyl)porphine (TCPP) and polyethylenimine (PEI), and subsequent surface modification with γ-poly(glutamic acid) (γ-PGA). TCPP served a dual function as the FL imaging probe and the photosensitizer. The as-prepared TCPP@PEI/PGA nanoparticles showed excellent water-solubility and biocompatibility, while having outstanding capabilities of in vivo bioimaging and 1O2 generation. FL bioimaging of mice and effective killing of CT 26 cells as well as CT 26 tumor-bearing mice upon laser irradiation were successfully demonstrated when using TCPP@PEI/PGA as theranostic nanoprobes. This study provides a simple but robust method to design and synthesize porphyrin-based polymer nanoparticles for theranostics.  相似文献   

18.
The Raman and infrared spectra of N-(2-pyridyl)thioformamide and N-(2-pyridyl)-thioacetamide have been measured. The assignment of the bands is aided by the complete normal coordinate treatment for all the vibrations of N-(2-pyridyl)thioformamide and its N-deuterated molecule using a Urey—Bradley force function for the in-plane vibrations and a valence force function for the out of plane vibrations. Variable temperature 1H NMR study of the two pyridylthionamides has also been performed. It is inferred that while N-(2-pyridyl)thioformamide favours a cis —CSNH— group, the other compound favours a trans —CSNH— grouping at ambient temperature.  相似文献   

19.
Upon treatment of ethyl 2-(4-chloro-2-pyridyl)benzoic acid, 2-(4-chloro-2-pyridyl)benzoate, and N,N-diisopropyl-2-(4-chloro-2-pyridyl)benzamide with LTMP at −75 °C in THF, the lithio derivatives at C5′ are generated regiospecifically, as demonstrated by subsequent quenching with electrophiles. The lithio derivative at C3′ is only evidenced from the benzamide at higher temperature (−50 °C), when treated with LTMP in THF; it instantly cyclizes to 1-chloro-4-azafluorenone. The latter is converted to onychine, an alkaloid endowed with anticandidal activity.  相似文献   

20.
The standard ( po =  0.1 MPa) molar enthalpies of combustion in oxygen, at T =  298.15 K, were measured by rotative bomb calorimetry for crystalline N, N -diethyl- N-furoylthiourea, (2-C4H3O)CONHCSN(C2H5)2, HFET, and N, N -diisobutyl- N-furoylthiourea, (2-C4H3O)CONHCSN(iso-C4H9)2, HFIB. The standard molar enthalpies of sublimation of both HFET and HFIB were measured by high-temperature Calvet microcalorimetry. These values were used to derive the standard molar enthalpies of formation of the compounds, in their crystalline and gaseous phases.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号